You are on page 1of 8

5th Flow Control Conference AIAA 2010-4708

28 June - 1 July 2010, Chicago, Illinois

Modification of the Laminar-to-Turbulent Transition on a


Flat Plate Using a DBD Plasma Actuator

Romain JOUSSOT1 and Dunpin HONG2


GREMI, CNRS/University of Orléans, 45076 ORLEANS Cedex 2, FRANCE

and

Régine WEBER-ROZENBAUM3 and Annie LEROY-CHESNEAU4


Institut PRISME, University of Orléans, 45072 ORLEANS Cedex 2, FRANCE

Active flow control by plasma actuators, usually called ElectroHydroDynamic actuators,


is currently under investigation in order to modify and control external flows. One of these
actuators consists in using a surface dielectric barrier discharge by creating a non-thermal
plasma at the dielectric surface. The plasma induces a low-velocity airflow, the “ionic wind”,
which adds momentum close to the wall. In this study, the ability of such an actuator to
delay or promote the transition of a boundary layer developed along a flat plate is studied
experimentally.

Nomenclature
δ1 = displacement thickness, mm
δ99 = boundary layer thickness, mm
f = frequency, Hz
H12 = shape factor, <>
Res = Reynolds number based on the length s, <>
s = curvilinear abscissa, mm
U = horizontal velocity component, m/s
Uact = maximum ionic wind expected without external flow, m/s
Ue = free-stream velocity, m/s
URMS = root mean square value of the wall-normal velocity, m/s
y = coordinate in wall-normal direction, mm

I. Introduction

A CTIVE flow control by plasma actuators, or ElectroHydroDynamic (EHD) actuators, is currently studied in
order to control external flows on aerodynamic geometries. Moreau1 and Corke2,3 give an overview of the
various configurations used and highlight the ability of this type of actuators to manipulate flows. In the last few
years, plasma actuators based on the surface Dielectric Barrier Discharge (DBD) have attracted increasing attention.
To generate such a discharge, a high voltage is applied to two metal electrodes asymmetrically placed on either side
of a dielectric material. The high voltage is generally a sine waveform and the non-thermal plasma produced on the
surface of the dielectric induces a flow called ionic wind. It is well accepted that the ionic wind results from the
transfer momentum between plasma ions and neutral molecules of the surrounding gas. Since the ionic wind is
induced close to the wall, the momentum is added directly inside the lower parts of the boundary layer. Depending
on the objectives of the control flow, results with different efficiencies have been obtained when the actuators were
operated in steady or unsteady (pulsed) mode.

1
PhD Student, GREMI, 14 Rue d’Issoudun, 45076 ORLEANS Cedex 2, Student AIAA Member.
2
Professor, GREMI, 14 Rue d’Issoudun, 45076 ORLEANS Cedex 2.
3
Associate Professor, iPRISME, 8, Rue Léonard de Vinci, 45072 ORLEANS Cedex 2.
4
Associate Professor, iPRISME, 8, Rue Léonard de Vinci, 45072 ORLEANS Cedex 2.
1
American Institute of Aeronautics and Astronautics

Copyright © 2010 by R. Joussot. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
Manipulating the laminar-to-turbulent transition is possible using plasma actuators. For low free-stream
velocities (less than 20 m/s), the transition has been successfully promoted or delayed with plasma actuators acting
on natural laminar boundary layers evolving on flat plates. Seraudie et al.4, Porter et al.5, and Magnier et al.6
promoted the transition, moving the transition point upstream of the natural position using DBD actuators working
in steady mode by amplifying disturbances in the boundary layer. In these studies, the aerodynamic conditions
(pressure gradient), the configuration of the actuator (power supply, geometry) and the location of the action
differed.
In order to achieve a delay of the laminar-to-turbulence transition, the main actuation parameters have to be
adjusted as closely as possible to the boundary layer state, since the location at which the transition occurs depends
on several factors such as surface roughness or free-stream turbulence for example. When the turbulence level is
low, the transition process is characterised by the development of growing instabilities, dominated in the initial
stages by 2D Tollmien-Schlichting waves, followed by 3D instabilities leading to turbulence. Different actuation
modes can be performed in order to delay the laminar-to-turbulence transition.
Grundmann and Tropea7 showed that two DBD actuators operated steadily in streamwise operation could
significantly reduce the perturbations which were introduced into the boundary layer by a third DBD actuator
operated unsteadily at an upstream position to promote transition. This actuation proved capable of modifying
boundary layer profiles to achieve more stable profiles. In Boucinha et al.8, a single plasma actuator was mounted on
a relatively thick flat plate to act steadily as a co-flow on the natural laminar evolving boundary layer. As the ionic
wind was fully contained within the laminar boundary layer (y < 3mm), a momentum addition and a decrease in the
boundary layer thickness were observed above and downstream of the actuator. In both these studies, actuation was
performed with the aim of modifying boundary layer profiles into more stable profiles by damping natural
instabilities, and hence, by reducing the velocity fluctuations.
Other studies deal with the control of specific instabilities occurring in the transition process. By means of a
DBD actuator operated in unsteady mode, Grundmann and Tropea9 observed a transition delay since they succeeded
in damping the amplitude of the velocity fluctuations at a given excitation frequency corresponding to artificially
excited Tollmien-Schlichting waves. Control of the transient growth in a Blasius boundary layer was recently
demonstrated experimentally by Hanson et al.10. Transient growth was induced using calibrated roughness elements,
and actuators consisted of a spanwise array of symmetric plasma actuators capable of generating spanwise-periodic
counter-rotating vortices. The aim was to apply a lower amplitude disturbance suitable for attenuating streamwise
streaks in the boundary layers, leading to a transition delay.
Following the results presented in Boucinha et al.8, in which a transition delay was observed by manipulating a
natural boundary layer evolving along a flat plate using a single plasma actuator, more experiments were performed
with the main objective of improving our understanding of the precise mechanisms of momentum addition to the
boundary layers. The results are discussed in this paper. Various locations of a single actuator and high voltage
amplitudes were tested to investigate the ability of the actuation to promote or delay the transition with regard to the
natural state of the boundary layer. In this study, a flat plate was placed inside the test section of an open-circuit
wind tunnel. The turbulence intensity of the wind tunnel
was around 0.3 % and the free-stream velocity was 20 m/s.
The actuator could be moved to study the influence of
position of the local momentum addition on the boundary
layer transition. Velocity measurements were made using a
single hot-wire probe and were performed behind the
actuator to determine the nature of the boundary layer.
The experimental setup, details of the measurement
techniques, design actuators, and the natural case and non-
manipulated boundary layer evolving on the flat plate are
presented in section II. In section III the results are
discussed.

II. Experimental Setup and Natural Boundary


Layer

A. Experimental setup Figure 1. Evolution of the mean turbulence


The flat plate used in this study was the same as the one intensity and the pressure gradient over the flat
used by Magnier et al.6 and Boucinha et al.8. The 1 m long plate.
and 30 mm thick flat plate was made of PVC and was
2
American Institute of Aeronautics and Astronautics
divided into three parts: the leading edge of a NACA 0015 profile (0-30% chord, 60 mm length), the plate on which
the actuators were mounted, and the trailing edge (30-100% chord) of the NACA0015 profile. In order to take into
account the leading edge curvature, the curvilinear abscissa s was used to determine the positions along the flat
plate. Experiments were performed in a subsonic open-circuit wind tunnel with a 2 m long square test section
(50 cm × 50 cm). The nozzle had a section contraction ratio of 1:16 resulting in a mean turbulence ratio of 0.3 % at
the beginning of the test section (Fig. 1). The 0.3 m wide plate was not directly positioned inside the test section of
the wind tunnel but was first placed between two vertical plates (1.5 m × 50 cm) and then inserted into the test
section. Measurements of the free-stream velocity were made 8 cm above the flat plate (Fig. 1) at 20 m/s. The
velocity remained more or less constant at the beginning of the plate where the actuators were placed, and a small
pressure gradient was observed from s = 250 mm.
Velocity profiles were performed using hot-wire anemometry. A 1-D hot-wire probe (Dantec® ref. 55P15) was
used to measure the longitudinal component of the velocity (i.e. the main flow direction). The hot-wire was
calibrated each day with the Dantec Dynamics Hot-Wire Calibrator (26 points between 0.5 and 30 m/s were taken).
The acquisitions were performed using the software StreamWare from Dantec®. Temperature corrections were
made to take variations in room temperature into account in the velocity measurements. The acquisition rate was
30 kHz and the signal was low-pass filtered at a cut-off frequency of 15 kHz. Each point was acquired during 11 s.
The probe was mounted on a 2-D traversing system (Isel® Automation LF5 series with a resolution step of 0.1 mm ±
0.02 mm). An optical measurement was performed to estimate the initial position of the hot-wire near the wall (the
precision was estimated at 0.05 mm).

a)

b)

Figure 2. Experimental setup. a) Schematic of the plasma actuator and b) positions of the actuators on the flat
plate.

B. EHD actuator and power supply


The EHD actuator consisted of two copper electrodes flush mounted on both sides of a multiple-layer dielectric.
The 6 mm-wide electrodes were separated by a 3 mm-wide gap (Fig. 2a). The upper electrode (86 µm thick) was
connected to a high voltage power supply (home-built, 0-40 kVPP, 0.1-10 kHz) and the second one was buried in
order to inhibit discharge at the grounded side. The flat plate can be equipped with 12 single actuators, and to
prevent any disturbance on the boundary layer only the working upper electrode was set up on the plate during
experiments. The first plasma actuator was placed at an abscissa of 117 mm from the leading edge of the plate and
the gap between actuators was 40 mm (Fig. 2b). The plasma actuators were oriented in order to induce an ionic wind
in the same direction as the free air stream.
It is well known that the mean flow created by an EHD actuator without an external flow can be compared to a
wall jet: the maximum velocity is very close to the wall. Boucinha11 has shown that this assumption is verified only
in the first 60 % of the plasma extension length. Behind this area the wall jet is diffuse: the maximum velocity
decreases and moves away from the wall. In the present paper, the frequency of the signal was fixed at 1 kHz and
3
American Institute of Aeronautics and Astronautics
the high voltage amplitude was 16, 20 or 24 kVPP. With these characteristics, the maximum velocities without an
external flow were 2.2, 3.3, and 4.4 m/s respectively. These velocities remain very low with regard to the external
flow velocity fixed at 20 m/s.
a) b)

Figure 3. Natural flow developing over the flat plate. a) Non-dimensional velocity profiles and (b) boundary
layer thickness as a function of the curvilinear abscissa.

C. Natural Case and Non-Manipulated Boundary Layer


First of all, the boundary layer was explored without any high-voltage electrode mounted on the dielectric. We
retrieve the natural evolution of a boundary layer over a flat plate, since the velocity profiles closely match the
Blasius and Prandtl profiles for laminar and turbulent
boundary layers, respectively. The position of the laminar-to-
turbulent transition was estimated by plotting the evolution of
the boundary layer thickness δ99 or the shape factor (Fig. 3). It
occurred around the abscissa 437 mm. This corresponds to a
Reynolds number Res of about 5.5 105. The position of the
transition in the natural flow is therefore in accordance with
the literature.
Only one high-voltage electrode was placed on the flat
plate in order to limit disturbances in the boundary layer. To
ensure the non-intrusivity of the electrode, the evolution of the
shape factor H12 along the plate was studied (Fig. 4). The
comparison with the natural case shows that the position of the
laminar-to-turbulent transition was only slightly affected by
the high-voltage electrode, even if it was placed near the
leading edge (the plasma begins at s = 117 mm for actuator 1). Figure 4. Shape factor as a function of the
The modification in the boundary layer evolution along the flat curvilinear abscissa with or without HV
plate cannot be explained by the presence of the high-voltage electrodes placed.
electrode.

III. Results on the Manipulated Boundary Layer


In this part, the results of only three positions of actuation (actuators 1 to 3) at three voltage amplitudes (16, 20,
and 24 kVPP) are presented. The frequency of the signal was fixed at 1 kHz and the plasma actuator worked in a
steady mode. Depending on the position and the signal voltage, both effects were observed on the boundary layer:
the laminar-to-turbulent transition was either delayed or promoted.

A. Modification of the position of the laminar-to-turbulent boundary layer


With our electrical parameters, the laminar-to-turbulent boundary layer transition was promoted only for
actuators 1 or 2 (Fig. 5a). For the first actuator (placed at s = 117 mm), the transition was promoted for the three
4
American Institute of Aeronautics and Astronautics
voltage amplitudes tested, occurring around 160 mm earlier than for the natural case. However, the action seems to
be less efficient (for promoting the transition) for the highest voltage tested. In order to obtain an accurate value of
the position of the transition, a larger number of profiles are required. For the second actuator (s = 157 mm), the
transition was promoted only for the lowest voltage tested (Fig. 5a). In comparison with the results obtained for
actuator 1, the transition occurred 80 mm upstream from the natural point. The action seems to be less efficient than
with actuator 1.
For actuator 2 and a voltage amplitude of 20 kVPP, and for actuator 3 at 16 kVPP, the laminar-to-turbulent
boundary layer transition seemed to be slightly delayed
a) (Fig. 5b). Many more profiles in the transition area are
necessary to quantify the transition delay, if indeed it
exists, in these cases. It will be considered that the plasma
has no effect in these two configurations.
For the last three configurations tested, the transition
was delayed (Fig. 5c). The comparison of the two
configurations with actuator 3 suggests that the action is
more efficient for a higher voltage, and so, for a higher
induced ionic wind. The transition delay was about
200 mm for the best configuration (actuator 3 at 24 kVPP).
For a given high-voltage amplitude (24 kVPP), the
transition was delayed both for actuator 2 and for actuator
3. For a lower voltage (20 kVPP), the delay occurred only
for actuator 3.
b)
Table 1. Action on the position of the laminar-to-
turbulent transition.
16 kVPP 20 kVPP 24 kVPP
Uact=2.2 m/s Uact=3.3 m/s Uact=4.4 m/s
Actuator 1
δ99 ~ 1.7 mm Promoting Promoting Promoting -
Reδ1 = 840
Actuator 2
δ99 ~ 1.8 mm Promoting - No effect Delaying
Reδ1 = 760
Actuator 3
δ99 ~ 2.0 mm No effect Delaying Delaying +
Reδ1 = 940

c) The efficiency of a plasma actuator is thus not only


due to the high voltage applied (i.e. the maximum ionic
wind velocity or the added momentum), but to the
position of the actuator on the plate (Table 1). For a given
actuator position, the transition delay seems to be greater
for higher added momentum. This added momentum has
to be lower when the actuator is placed closer to the
natural transition point. It should be remembered that
these results were obtained for a given signal frequency
and a given free-stream velocity. The boundary layer
thickness, the ratio between the maximum ionic wind Uact
and the free-stream velocity Ue, and the stability of
natural perturbations will be affected if the frequency and
the free-stream velocity change. The values required to
Figure 5. Shape factor as a function of the promote or delay the laminar-to-turbulent transition will
curvilinear abscissa. The transition is a) promoted, then vary from those presented here, but the tendencies
b) not moved, and c) delayed. will probably be the same.

5
American Institute of Aeronautics and Astronautics
B. Modification of the velocity profiles
Hot-wire measurements were performed far from the plasma zone to avoid electrical arcs. The first profiles were
made 4 cm after the end of the plasma zone. Furthermore the ratio between the ionic wind and the free-stream
velocity was low. It was observed that the mean velocity profiles measured do not appear to be modified when the
actuator is operating. LDA measurements in the vicinity of the plasma zone would be necessary to observe the
modification of velocity profiles. However, even far from the plasma zone, the RMS profiles were modified when
the actuator was operating (Fig. 6).

Figure 6. Evolution of the URMS profiles along the flat plate. Plasma off in blue circles, actuator 2 acting at
16 kVPP in red triangles and at 24 kVPP in green diamonds.

The natural evolution of the URMS without plasma is similar to results published in the literature (in blue on Fig.
6). The URMS levels are stable as the boundary layer is laminar. Near the transition zone, the maximum of URMS
grows slowly to reach its maximum at the transition point (around s = 437 mm). When the boundary layer is
turbulent, the levels decrease and become stable at relatively high values (around 0.08Ue near the plate). Tollmien-
Schlichting waves are visible in the URMS profiles. For the actuation at 24 kVPP (in green on Fig. 6), the maximum of
URMS is twice as great as in the natural case. Nevertheless, this level remains constant much longer. The velocity
profiles are thus stabilized. When the transition occurs (around s = 597 mm), the evolution is classical. As shown for
the shape factor (Fig. 5), this corresponds well to a delay of the laminar-to-turbulent transition. For the actuation at
16 kVPP (in red on Fig. 6), the evolution of the URMS profiles is comparable to the natural case but occurs earlier. The
transition is then slightly promoted as shown by the shape factor between s = 357 mm and s = 437 mm.

Figure 7 shows the power spectra densities of the velocity fluctuations measured 5 cm downstream of the
beginning of the plasma zone at a height of 0.2 mm above the plate when the actuator 2 is operating. Here we
present only the results obtained at a height of one of the most effective dampings, which is the height of the greatest
energy fluctuations. Indeed, the effect of the actuator on the natural instabilities was observed to be dependent on the
height, as described by Grundmann and Tropea9.
All the spectra show the peak at the frequency of f = 1 kHz, which corresponds to the signal frequency. In fact,
the plasma discharge couples momentum to the ambient air with a linear frequency forcing which is the same as the
frequency of the power supply of the actuator8.
6
American Institute of Aeronautics and Astronautics
Figure 7. Power density spectrum of velocity fluctuations 4 cm downstream of the actuator. Plasma off in
red, actuator 2 acting at 16 kVPP in green, and at 24 kVPP in blue. Downstream of actuator a) 1, b) 2, and c) 3.

For the natural (in red) and the manipulated boundary layer, two main peaks at a frequency of about 55 and
110 Hz were found. They correspond to the Tollmien-Schlichting waves. When one actuator is working at any HV
amplitude, their amplitudes are always higher than with the actuator turned off. However, since the energy is
contained within a frequency bandwidth that is higher when the actuator is working, their relative amplitudes appear
lower. When the transition is delayed (actuator 2 at 24 kVPP for example), both peaks are strongly reduced. Enough
momentum is added in the boundary layer to mix the TS-waves in a high level of energy in order to reduce their
relative amplitude and thus to delay their natural amplification. The instabilities are sufficiently damped and the
transition is delayed.

IV. Conclusion
This paper deals with the modification of the laminar-to-turbulent transition on a flat plate using a single DBD
plasma actuator steadily operated with an AC wave form. Previous work has already shown that this actuation mode
can significantly reduce natural instabilities by stabilizing the velocity profiles of the boundary layer, since such a
plasma discharge couples momentum to the ambient air in the vicinity of the plate. This study is expected to aid in
the understanding of the effects of momentum addition to the boundary layer. Focusing on the effects of the position
of the actuator and the high voltage amplitude, namely the ratio between the maximum ionic wind Uact and the free-
stream velocity Ue, results show that it is possible to promote or to delay the transition. Finally, in order to achieve
the transition delay, this study highlights that the main actuation parameters have to be adjusted as closely as
possible to the boundary layer state to delay the natural amplification of the Tollmien-Schlichting waves by
stabilizing the velocity profiles of the boundary layer. This work will shortly be completed by the study of the high
voltage frequency and a linear stability analysis to gain a deeper understanding into this actuation mode.

References
1
Moreau, E., “Airflow control by non thermal plasma actuators”, Journal of Physics D : Applied Physics,Vol. 40, 2007, pp
605-636.
2
Corke, T.C., Post, M.L., and Orlov, D.M., “Single dielectric barrier discharge plasma enhanced aerodynamics: physics,
modelling and applications”, Experiments in Fluids, Vol. 46, No. 1, 2009, pp1-26.
3
Corke, T.C., Enloe, C.L., and Wilkinson S.P., “Dielectric Barrier Discharge plasma actuators for flow control”, Annual
Review of Fluid Mechanics, vol. 42, 2010, pp. 505-529.
4
Seraudie, A., Aubert, E., Naude, N., and Cambronne, J., “Effect of plasma actuators on a flat plate laminar boundary layer
in subsonic conditions”, AIAA Paper n°2006-3350, San Fransisco, 2006.
5
Porter, C. O., McLaughlin, T. E., Enloe, C. L., Font, G. I., Roney, J., and Baughn, J. W., “Boundary layer control using
plasma actuator”, AIAA Paper n°2007-786, Reno, 2007.
6
Magnier, P., Boucinha, V., Dong, B., Weber, R., Leroy-Chesneau, A., Hong, D., and Hureau, J., “Experimental study of the
flow induced by a sinusoidal dielectric barrier discharge actuator and its effects on a flat plate natural boundary layer”, Journal
Fluids Engineering, Vol. 131, No. 1, 2009, 011203 (11pp).
7
Grundmann, S. and Tropea, C., “Experimental transition delay using glow-discharge plasma actuators”, Experiments in
Fluids, Vol. 42, No. 4, 2007, pp. 653-657.
8
Boucinha, V., Magnier, P., Leroy-Chesneau, A., Weber, R., Joussot, R., Dong, B., and Hong, D., “Characterization of the
ionic wind induced by a sine DBD actuator used for laminar-to-turbulent transition delay”, AIAA Paper n°2008-4210, Seattle,
2008.

7
American Institute of Aeronautics and Astronautics
9
Grundmann, S. and Tropea, C., “Active cancellation of artificially introduced Tollmien-Schlichting waves using plasma
actuators”, Experiments in Fluids, Vol. 44, No 5, 2007, pp 795-806.
10
Hanson, R.E., Lavoie, P., Naguib, A.M., and Morrison, J.F., “Transient growth instability cancelation by a plasma actuator
array”, Experiments in Fluids, 2010, DOI 10.1007/s00348-010-0877-1.
11
Boucinha, V., “Etude de l’écoulement induit par une décharge à barrière diélectrique surfacique – Contribution au contrôle
des écoulements subsoniques par actionneurs plasmas”, PhD Dissertation, University of Orléans, 2009.

8
American Institute of Aeronautics and Astronautics

You might also like